Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access February 10, 2021

Recovery of γ-Fe2O3 from copper ore tailings by magnetization roasting and magnetic separation

  • Bing Luo , Tongjiang Peng EMAIL logo and Hongjuan Sun EMAIL logo
From the journal Open Chemistry

Abstract

To comprehensively reuse copper ore tailings, the recovery of γ-Fe2O3 from magnetic roasted slag after sulfur release from copper ore tailings followed by magnetic separation is performed. In this work, after analysis of chemical composition and mineralogical phase composition, the effects of parameters in both magnetization roasting and magnetic separation process with respect to roasting temperature, residence time, airflow, particle size distribution, magnetic field intensity, and the ratio of sodium dodecyl sulfonate to roasted slag were investigated. Under optimum parameters, a great number of γ-Fe2O3 is recycled with a grade of 66.86% and a yield rate of 67.21%. Meanwhile, the microstructure, phase transformation and magnetic property of copper ore tailings, roasted slag, and magnetic concentrate are carried out.

1 Introduction

With urgent requirement of the thirteenth 5-Year Plan for circular economy in China, copper ore tailings as one kind of non-ferrous tailings in industrial solid waste have been paid more and more attention to comprehensive disposal and resource utilization. It is estimated that approximately 400 tons of tailings are generated, while every ton of copper produced [1]. Copper ore tailings contain pyrrhotite (Fe(1–x)S, where 0 ≤ x < 0.223), which has three homogeneous polychromatic variants of iron sulfide mineral and tends to produce acid mine drainage because of oxidation [2,3]. To effectively prevent the environmental risk of the copper ore tailings and recycle the sulfur resource in the copper ore tailings, the sulfur released as SO2 can be collected to produce sulfuric acid by oxidation roasting as mentioned in the previous study [4]. Meanwhile, the iron units are concentrated in the roasted slag and existed in the form of γ-Fe2O3. Maghemite is a defective spinel structure where Fe3+ occupies both of A tetrahedral position and of B tetrahedral position with a certain amount of vacancies [5,6]. In addition, γ-Fe2O3 is widely used as a magnetic material with advantages such as stabilization, nontoxicity, and low consumption, which has a wide range of application prospects [7]. Therefore, the recovery and utilization of γ-Fe2O3 in roasted copper ore tailings are of great importance to save iron resources and promote copper ore tailings more sustainable and respectful to ecological environment.

As we all know, both iron concentrate (Fe3O4) and sponge iron (Fe) have some magnetic properties; therefore, magnetic separation is an effective way to recover iron where Fe3O4 and Fe are used for iron smelting and steel making [8,9,10,11], respectively. Currently, the recovery of Fe3O4 and Fe was studied by many researchers using magnetic separation from various metal dressing tailings or other solid wastes rich in iron [12,13,14]. However, not all iron with a certain phase in tailings and solid wastes are Fe3O4 or Fe. It requires magnetizing roasting and direct reduction to modify nonmagnetic iron to magnetic iron [15,16,17,18]. The modified conditions with iron-bearing minerals are summarized in Table 1. However, the current studies about the recovery of γ-Fe2O3 remain unsolved. Taking advantage of magnetic properties of γ-Fe2O3 will greatly facilitate the recovery of iron in the roasted slag.

Table 1

The modified conditions with iron-bearing minerals

Iron-bearing minerals Modified condition Recovery of iron
Fe2SiO4 With weak oxidation modification roasting in CO–CO2 mixed gases, FeO in Fe2SiO4 was oxidized to Fe3O4 Grinding and magnetic separation
Fe2O3 Under high temperature (usually >1,000°C), it was reduced to Fe3O4 by reducing agent (pulverized coal) Magnetic separation
Fe(1–x)S In coal-based direct reduction with non-coking coal, the final product was Fe, which was used for steelmaking and as a charge of blast furnaces Magnetic separation

Therefore, sulfur release was studied first in previous study [4]. In this paper, the magnetizing roasting of the copper ore tailings followed by magnetic separation is further investigated to recover γ-Fe2O3 from the roasted slag. The roles of roasting temperature, residence time, airflow, particle size, magnetic field intensity, and the ratio of sodium dodecyl sulfonate (SDS) to roasted slag were established. Furthermore, the characteristics of magnetic concentrate were discussed as well. More usefully, it born the potential to comprehensively use copper ore tailings and reduce environmental risks.

2 Materials and methods

2.1 Materials

Copper ore tailings were sampled from Liwu copper of Jiulong County in Ganzi Prefecture of Sichuan Province, China. Copper ore tailings were dried at 105 ± 2°C for 8 h, milled to 100 meshes, and stored, sequentially. The mineralogical phase compositions of the copper ore tailings were determined by X-ray diffraction (XRD, The Netherlands), and the results are shown in Figure 1. The chemical composition of the copper ore tailings was tested by X-ray fluorescence (XRF, The Netherlands), and the results are shown in Table 2.

Figure 1 
                  XRD pattern of the copper ore tailings sample.
Figure 1

XRD pattern of the copper ore tailings sample.

Table 2

Chemical composition of the copper ore tailings (wt%)

Components Fe2O3 SO3 SiO2 Al2O3 Na2O ZnO CaO K2O MgO
Content 36.35 27.82 22.81 6.03 2.12 1.61 1.16 0.93 0.47
Components TiO2 CuO P2O5 Co3O4 PbO ZrO2 MnO SrO Total
Content 0.25 0.22 0.10 0.05 0.03 0.02 0.02 0.01 100

The main reagent SDS was chemically pure, which was bought from Nanjing Chemical Reagent Co Ltd, China.

The main instruments were LTKC-6-16A programmable energy-saving tubular furnace (Hangzhou Blue Sky Instrument, Co Ltd, China) and XCRQ-50*70 magnetic tube (Wuhan prospecting machinery Factory, Co Ltd, China).

2.2 Experimental procedures

Six group experiments were carried out to optimize different parameters in both magnetization roasting and magnetic separation on γ-Fe2O3 recovery from copper ore tailings, including the ratio of coal to iron ore tailings, roasting temperature, roasting time, airflow, particle size, magnetic field, and the ratio of SDS. The sample was located in a programmable energy-saving tubular furnace (LTKC-6-16A) to magnetizing roast. The magnetized sample was weighed about 20 g in low-intensity magnetic separation first. Subsequently, the best parameters of magnetizing roast were confirmed. Under the optimized parameters, 100 g magnetized sample and 4 g SDS were mixed evenly and then transferred into a 1,000 mL beaker with about 500 mL water and an appropriate amount of alcohol. The magnetic mixture was stirred well and allowed to stand for 10 min. Subsequently, the magnetic mixture was put in the magnetic tube (XCRQ-50*70), in which the optimized magnetic field intensity was investigated by changing current intensity.

2.3 Analytical methods

The mineralogical phase compositions of both raw tailings and roasted samples were tested by X-ray diffraction (XRD, X’Pert pro + 3 kW, The Netherlands, λ = 0.15406 nm), using Cu Kα radiation with continuous scanning step of 0.02° in 2θ ranging from 3 to 80°. The chemical composition of the copper ore tailings was obtained by X-ray fluorescence Spectrometer (XRF- Axios, The Netherlands, wavelength dispersive type, 2.4 kW) while the content ranges from 0.01 to 100% and the sample diameter is 32 mm, using Rh target with θ/2θ scanning mode of the goniometer according to JY/T 016-1996. The magnetic properties of magnetic concentrate and copper ore tailings were investigated using a vibrating sample magnetometer (VSM BKT-4500Z, China). The microstructures and elemental analysis of copper ore tailings, roasted slag, and magnetic concentrate were characterized by scanning electron microscopy (Leica Cambridge LTD, 15 kV), with magnifying multiple ranges from 3,000 to 200. The content of γ-Fe2O3 was determined by chemical titration. The grade of γ-Fe2O3 was equal to the content of γ-Fe2O3.

The yield of γ-Fe2O3 was calculated as follows:

(1) i = Q j Q y × 100 % ,

where i is the yield of γ-Fe2O3; Q j is the magnetic concentrate mass; and Q y is the roasted slag mass.

  1. Ethical approval: The research conducted is not related to either human or animal use.

3 Results and discussion

3.1 Characterization of the copper ore tailings

From Figure 1, the main metal phase in samples was pyrrhotite and lepidocrocite. After pyrrhotite and lepidocrocite were converted to iron in certain ways [12,14,16], iron could be recycled as an important metal resource for further use. As shown in Table 2, the content of Fe2O3 reached 36.35 wt%. When the copper ore tailings were roasted from 600 to 1,200°C, the roasted slag was magnetic which contained γ-Fe2O3 [4]. Therefore, it was necessary and meaningful to roast to release sulfur first and recycle γ-Fe2O3 subsequently.

3.2 γ-Fe2O3 recovery

3.2.1 Effect of the roasting temperature on γ-Fe2O3 recovery

Roasting temperature was the most important factor that affects sulfur release and magnetization. Sulfur compounds were easy to decompose and release sulfur with the increasing roasting temperature ranging from 1,000 to 1,350°C [19]. After roasting at 1,200°C with the duration time from 20 to 60 min, the sulfur release rate reached to 99.78% and the residual S content reduced to 0.06% [4]. In this study, roasting temperature was set as 1,000, 1,050, 1,100, 1,150, and 1,200°C, and roasted time was 10 min, with the airflow of 5 L min−1. The magnetic field intensity was 160 mT. Results are shown in Figure 2.

Figure 2 
                     Effect of roasting temperature on γ-Fe2O3 recovery.
Figure 2

Effect of roasting temperature on γ-Fe2O3 recovery.

As shown in Figure 2, with the roasting temperature increasing, the grade and yield of γ-Fe2O3 increased from 40.15 to 45.37% and from 7.69 to 52.06%, respectively. From 1,000 to 1,200°C, the grade of γ-Fe2O3 increased with the increase in roasting temperature. The reaction between Fe(1–x)S and O2 was dramatic because of sulfur release [20]. While at 1,050°C, the reaction was accelerated suddenly, which promotes the oxidation of Fe3O4 to γ-Fe2O3. When high roasting temperatures exceeded 1,200°C, the sintered sample would melt. Then, it was unfavorable to separate magnetic particle from the roasted slag.

3.2.2 Effect of residence time on γ-Fe2O3 recovery

Copper ore tailings were roasted at 1,200°C for different roasting time, with the airflow of 5 L min−1. The magnetic field intensity was 160 mT. Results are shown in Figure 3. From Figure 3, neither the grade nor the yield was in linear relationship with roasting time. After 10 min, the yield of γ-Fe2O3 increased even more and achieved a peak of 73.52%, whereas the grade of γ-Fe2O3 was slow, only 49.96%. It indicated that some non-magnetic particles, which attached to magnetic concentrate, were separated together in low-intensity magnetic separation [21].

Figure 3 
                     Effect of roasting time on γ-Fe2O3 recovery.
Figure 3

Effect of roasting time on γ-Fe2O3 recovery.

3.2.3 Effect of airflow on γ-Fe2O3 recovery

As an oxidant, air not only affected the rate of sulfur release but also had an influence on the magnetization roasting [22]. Copper ore tailings were roasted at 1,200°C for 30 min with different airflow, and the magnetic field intensity was 160 mT. As shown in Figure 4, whether the airflow was increased or decreased, both the grade and the yield of γ-Fe2O3 decreased. When airflow was low, iron atoms released from pyrrhotite lattice could not combine with oxygen atom completely because of sulfur atom undergoing external diffusion and desorbing as SO x [23]. When airflow was high, the excess air may decrease the roasting temperature and take away some powder with iron [24]. In this experiment, the optimum airflow was confirmed 5 L min−1.

Figure 4 
                     Effect of airflow on γ-Fe2O3 recovery.
Figure 4

Effect of airflow on γ-Fe2O3 recovery.

3.2.4 Effect of particle size distribution on γ-Fe2O3 recovery

Copper ore tailings were roasted at 1,200°C for 30 min, with the airflow of 5 L min−1. According to the approximate distribution of particle size in copper ore tailings, standard sieves were selected as 100, 120, 140, 160, and 180 meshes, respectively. The roasted samples were milled and sifted in turn. The magnetic field intensity was 160 mT. From Figure 5, the grade and the yield of γ-Fe2O3 increased to 51.20 and 76.25%, respectively, whereas the particle size was 120 µm. As a result of complex embedded features of roasted samples, magnetic separation was not carried out effectively with improper particle [25]. When the particle size was large, the non-magnetic particle would be attached to the magnetic particle and would be magnetized [26], which lead to the decrease in the grade and the increase in the yield simultaneously.

Figure 5 
                     Effect of size of particles on γ-Fe2O3 recovery.
Figure 5

Effect of size of particles on γ-Fe2O3 recovery.

3.2.5 Effect of magnetic field intensity on γ-Fe2O3 recovery

Copper ore tailings were roasted under optimal conditions and milled to 120 meshes. Subsequently, the samples were separated in different magnetic intensities selected as 160, 180, 200, 220, 240, and 250 mT, respectively. As shown in Figure 6, the grade and yield increased with magnetic intensity ranging from 160 to 220 mT. Although the magnetic intensity was low, magnetic particles were easily carried away by flowing water under the effects of gravity and scouring, which resulted in the loss of magnetic particles and decrease in grade and yield of γ-Fe2O3. Meanwhile, although the magnetic intensity was excessive, gangue particles were absorbed on magnetic particles and were separated together [27,28]. Therefore, the grade and yield of γ-Fe2O3 increased to 63.59 and 78.56% with the magnetic intensity 220 mT, respectively.

Figure 6 
                     Effect of magnetic field intensity on γ-Fe2O3 recovery.
Figure 6

Effect of magnetic field intensity on γ-Fe2O3 recovery.

3.2.6 Effect of the ratio of SDS to roasted slag on γ-Fe2O3 recovery

Because of the aggregation between magnetic and non-magnetic particles, the yield was always higher than the grade from the above experimental results. To separate magnetic and non-magnetic particles before magnetic separation, as a kind of surfactant, SDS could be used to disperse the magnetic and non-magnetic particles with certain levels [29]. The magnetic field intensity was 220 mT, with the proportion of SDS and samples at 1:100, 1:50, 3:100, 1:25, and 1:20. Results are shown in Figure 7. It revealed that the grade was increased and the yield was decreased with increasing SDS. When the ratio of SDS to roasted slag elevated from 1:100 to 1:25, the yield decreased dramatically from 72.63 to 67.21% with a corresponding grade ranging from 64.18 to 66.86%. The results showed that SDS could promote to strip and separate magnetic and non-magnetic particles. Therefore, the optimum ratio of SDS to roasted slag was suggested at 1:25.

Figure 7 
                     Effect of the ratio of SDS to roasted slag on γ-Fe2O3 recovery.
Figure 7

Effect of the ratio of SDS to roasted slag on γ-Fe2O3 recovery.

3.3 Analysis of roasted slag

3.3.1 Phase transformation

Owing to the characteristics of pyrrhotite with more sulfur and less iron, the purpose of magnetization roasting was to release sulfur first [4] and then to oxidize magnetite to maghemite. The possible reactions during magnetization roasting were expressed as follows:

(2) 4 Fe 1 x S + 7 O 2 2 ( 1 x ) Fe 2 O 3 + 4 SO 2 ,

(3) 3 Fe 1 x S + ( 5 2 x ) O 2 ( 1 x ) Fe 3 O 4 + 3 SO 2 ,

(4) 4 Fe 3 O 4 + O 2 6 γ-Fe 2 O 3 .

Equation (1) occurs below 1,000°C. Because of the presence of sulfur, the performance of iron products would be affected. Therefore, it was necessary to raise the temperature to release the sulfur before the recovery of iron. When the temperature reached 1,000°C, magnetite appeared as shown in Figure 8. With the increasing temperature, Fe3O4 was oxidized to γ-Fe2O3.

Figure 8 
                     XRD spectra of roasted slag at different roasting temperature.
Figure 8

XRD spectra of roasted slag at different roasting temperature.

To investigate the transformation of iron, the roasted slag with the roasting temperature from 1,000 to 1,200°C for 30 min at airflow 5 L min−1 was sampled and analyzed for mineral phase. From Figure 8, Fe3O4 appeared which was consistent with equation (2) and could also make the low-intensity magnetic separation easy to realize. Meanwhile, compared with the XRD spectra of copper ore tailings, the pattern of pyrrhotite disappeared, which indicated that pyrrhotite decompounds and was oxidized. Moreover, the pattern of Fe2O3 also showed up, which was corresponding to equation (1).

3.3.2 SEM-EDS analysis

The SEM images of copper ore tailings and roasted slag (1,000°C for 30 min, airflow 5 L min−1) are shown in Figure 9. Meanwhile, the energy spectrum of A/B area (Figure 9a and b) was also presented together. From Figure 9b, the roasted slag was light sintered. In the roasting process, crystal water of samples evaporated gradually and SO2 generated during the reaction volatilize [26]. This may have resulted in the appearance of porous and the increase in granularity [16]. Therefore, it was necessary and significant to grind and mill the sample to a certain extent. From the energy spectrum of roasted slag, S was still present in samples, which indicated that the sulfur release reaction was not complete. Therefore, the roasting temperature should be raised appropriately to release residual sulfur before the recovery of magnetic concentrate.

Figure 9 
                     The SEM images and energy spectrum of A/B area in copper ore tailings and roasted slag. (a) Copper ore tailings; (b) roasted slag at 1,000°C for 30 min, airflow 5 L min−1.
Figure 9

The SEM images and energy spectrum of A/B area in copper ore tailings and roasted slag. (a) Copper ore tailings; (b) roasted slag at 1,000°C for 30 min, airflow 5 L min−1.

3.4 Analysis of magnetic concentrate

3.4.1 XRD analysis

The magnetic concentrate was separated from the roasted slag as above, and Figure 10 presents its XRD pattern. From Figure 10, maghemite was the major metallic mineral phases in magnetic concentrate. Moreover, quartz fine-disseminated with maghemite was separated together. According to the report, magnetic concentrate could be sold directly when the content of total Fe and S exceed 65% and was less than 0.1% [21], respectively. Obviously, the content of γ-Fe2O3 was 66.86% with sulfur 0.08% from Table 3. Therefore, the magnetic concentrate powder could be used in the powder metallurgy industry.

Figure 10 
                     XRD spectra of magnetic concentrate.
Figure 10

XRD spectra of magnetic concentrate.

Table 3

Chemical composition of the magnetic concentrate (wt%)

Components Fe2O3 SiO2 Al2O3 MgO ZnO CaO Na2O K2O CuO Co3O4
Content 66.86 17.30 7.08 2.37 1.76 1.14 1.07 0.92 0.92 0.37
Components TiO2 P2O5 SO3 PbO BaO MnO ZrO2 Rb2O As2O3 Total
Content 0.30 0.12 0.08 0.04 0.03 0.03 0.02 0.01 0.01 100%

3.4.2 SEM-EDS analysis

Figure 11 shows the SEM image of magnetic and the energy spectrum of C area. From Figure 11, some magnetic concentrate were spherical and granular. Meanwhile, coarser particle in magnetic fraction and finer sized particles at the non-magnetic aggregated together. Fe, O, and Si were included in C area, which was corresponding to the XRD analysis.

Figure 11 
                     The SEM images and energy spectrum of C area in magnetic concentrate.
Figure 11

The SEM images and energy spectrum of C area in magnetic concentrate.

3.4.3 Magnetic property

The saturation magnetization (Ms) of copper ore tailings, roasted slag (1,200°C for 30 min, airflow 5 L min−1), and magnetic concentrate at above optimum conditions were studied in VSM, and the results are shown in Figure 12. It was obvious that Ms of copper ore tailings increases from 4.85 to 11.55 emu/g after magnetizing roasting, which indicated that the magnetic properties may vary more or less under different roasting conditions [30]. Moreover, the maximum Ms of the magnetic concentrate was 12.94 emu/g. These results were attributed to the conversion of weak Fe(1–x)S to maghemite in magnetizing roasting and the segregation of magnetic particles and non-magnetic particles by wet magnetic separation [31].

Figure 12 
                     Magnetization curve of copper ore tailings, roasted slag, and magnetic concentrate.
Figure 12

Magnetization curve of copper ore tailings, roasted slag, and magnetic concentrate.

3.5 The process steps in separation and recovery of γ-Fe2O3

The process of separation and recovery of γ-Fe2O3 is summarized in Figure 13. The pyrrhotite rich in copper ore tailings was oxidized to γ-Fe2O3 which was highly magnetic in magnetizing roasting. Meanwhile, sulfur in the form of SO2 was released from the inside of pyrrhotite to the surface through slits. After magnetizing roasting, magnetic particles and non-magnetic particles were coated with each other. Therefore, it was imperative to crush and mill roasted slag to disperse γ-Fe2O3 and Si–Ca–Mg–Al minerals. Subsequently, magnetic γ-Fe2O3 was separated in magnetic separation tube.

Figure 13 
                  The diagram of the process steps in separation and recovery of γ-Fe2O3 from copper ore tailings.
Figure 13

The diagram of the process steps in separation and recovery of γ-Fe2O3 from copper ore tailings.

4 Conclusions

  1. Major mineral phases of copper ore tailings were quartz, biotite, clinochlore, pyrrhotite, and lepidocrocite. The content of total Fe was 36.35%. Most of the iron existed in non-magnetic characteristics, which were hard to separate.

  2. With six groups’ experiments, the optimal roasting and magnetization separation conditions were determined as follows: roasting at 1,200°C for 30 min with airflow 5 L min−1 and the ratio of SDS to roasted slag (120 µm) is 1:25 in magnetic field intensity 220 mT. Under these conditions, the yield was 67.21%, the content of γ-Fe2O3 separated from roasted slag was 66.86% with sulfur 0.08%, and the Ms of magnetic concentrate was 12.94 emu/g.

  3. As for tailings or waste rich in sulfur and iron, the recycle process was demonstrated to be feasible, where sulfur release from copper ore tailings with optimal roasting conditions first and iron recovery from roasted slag subsequently.

  4. The technology of magnetization roasting and wet magnetic separation to recycle γ-Fe2O3 could be extended to other iron-bearing tailings or waste, including iron ore tailings, pyrite-containing tailings, pyrrhotite-containing tailings, and so on.

Acknowledgments

The authors would like to thank Long Shan academic research support program (No. 14tdgk02) and the project of Key research platform focused on teachers’ research innovation team (No. 17LZXT11) both from Southwest University of Science and Technology.

  1. Funding information: Funding was provided by Long Shan academic research support program (No. 14tdgk02) and the project of Key research platform focused on teachers’ research innovation team (No. 17LZXT11) both from Southwest University of Science and Technology.

  2. Authors’ contributions: B. Luo – methodology, writing original draft, preparation writing, reviewing, and editing; T. J. Peng – supervision and funding acquisition; H. J. Sun – software and data curation validation.

  3. Data availability statement: All data generated or analyzed during this study are included in this published article.

  4. Declaration of interests: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. The authors declare the following financial interests/personal relationships which may be considered as potential competing interests.

References

[1] Liu YJ, Huang LB. Magnetite recovery from copper tailings increases arsenic distribution in solution phase and uptake in native grass. J Environ Manage. 2016;186(SI2):175–82. 10.1016/j.jenvman.2016.05.025.Search in Google Scholar PubMed

[2] Manceau A, Merkulova M, Murdzek M, Batanova V, Baran R, Glatzel P, et al. Chemical forms of mercury in pyrite: implications for predicting mercury releases in acid mine drainage settings. Environ Sci Technol. 2018;52(18):10286–96. 10.1021/acs.est.8b02027.Search in Google Scholar PubMed

[3] Nleya Y, Simate GS, Ndlovu S. Sustainability assessment of the recovery and utilisation of acid from acid mine drainage. J Clean Prod. 2016;113:17–27. 10.1016/j.jclepro.2015.11.005.Search in Google Scholar

[4] Luo B, Peng TJ, Sun HJ, Ding WJ. Innovative methodology for sulfur release from copper tailings by the oxidation roasting process. J Chem-NY. 2020;2020:1–11. 10.1155/2020/8090846.Search in Google Scholar

[5] Saravanan P, Alam S, Mathur GN. Comparative study on the synthesis of γ-Fe2O3 and Fe3O4 nanocrystals using high-temperature solution-phase technique. J Mater Sci Lett. 2003;22:1283–5. 10.1023/A:1025470405000.Search in Google Scholar

[6] Wu W, Xiao XH, Zhang SF. Synthesis and magnetic properties of maghemite (γ-Fe2O3) short-nanotubes. Nanoscale Res Lett. 2010;5(9):1474–9. 10.1007/s11671-010-9664-4.Search in Google Scholar PubMed PubMed Central

[7] Aliahmad M, Moghaddam NN. Synthesis of maghemite (gamma-Fe2O3) nanoparticles by thermal-decomposition of magnetite (Fe3O4) nanoparticles. Mater Sci-Pol. 2013;31(2):264–8. 10.2478/s13536-012-0100-6.Search in Google Scholar

[8] Wu NT, Tang YP, Zeng M. Design of hybrid inorganic-organic nanosensor based on Fe3O4 as the core and recovery features. J Lumin. 2018;202:502–7. 10.1016/j.jlumin.2018.06.018.Search in Google Scholar

[9] Tu YJ, Lo SC, You CF. Selective and fast recovery of neodymium from seawater by magnetic iron oxide Fe3O4. Chem Eng J. 2015;262:966–72. 10.1016/j.cej.2014.10.025.Search in Google Scholar

[10] Geng C, Wang HJ, Hu WT, Li L, Shi CS. Recovery of iron and copper from copper tailings by coal-based direct reduction and magnetic separation. J Iron Steel Res Int. 2017;24(10):991–7. 10.1016/S1006-706X(17)30145-0.Search in Google Scholar

[11] Zhang XK, Zhou KG, Chen W. Recovery of iron and rare earth elements from red mud through an acid leaching-stepwise extraction approach. J Cent South Univ. 2019;26(2SI):458–66. 10.1007/s11771-019-4018-6.Search in Google Scholar

[12] Zhang XZ, Zhao G, Cao Y. Magnetothermal heating facilitates the cryogenic recovery of stem cell-laden alginate-Fe3O4 nanocomposite hydrogels. Biomater Sci-UK. 2018;6(12):3139–51. 10.1039/ c8bm01004h.Search in Google Scholar

[13] Gubin AS, Sukhanov PT, Kushnir AA, Proskuryakova ED. Recovery and preconcentration of phenols from aqueous solutions with a magnetic sorbent based on Fe3O4 nanoparticles and hyper-cross-linked polystyrene. Russ J Appl Chem. 2018;91(10):1626–34. 10.1134/S1070427218100099.Search in Google Scholar

[14] Yu JW, Han YX, Li LJ. Separation and recovery of iron from a low-grade carbonate-bearing iron ore using magnetizing roasting followed by magnetic separation. Sep Sci Technol. 2017;52(10):1768–74. 10.1080/01496395.2017.1296867.Search in Google Scholar

[15] Ma NY, Houser JB, Wood LA. Enhancement of iron recovery from steelmaking slag fines by process optimization of upgrading the slag fines with dry magnetic separation. J Sustain Metall. 2017;3(2):280–8. 10.1007/s40831-016-0079-z.Search in Google Scholar

[16] Peek E, Barnes A, Tuzun A. Nickeliferous pyrrhotite – “waste or resource”? Min Eng. 2011;24(7SI):625–37. 10.1016/j.mineng.2010.10.004.Search in Google Scholar

[17] Sui YL, Guo YF, Jiang T, Qiu GZ. Separation and recovery of iron and titanium from oxidized vanadium titano-magnetite by gas-based reduction roasting and magnetic separation. J Mater Res Technol. 2019;8(3):3036–43. 10.1016/j.jmrt.2018.05.031.Search in Google Scholar

[18] Min XB, Zhou BS, Ke Y. Sulfidation behavior of ZnFe2O4 roasted with pyrite: sulfur inducing and sulfur-oxygen interface exchange mechanism. Appl Surf Sci. 2016;371:67–73. 10.1016/j.apsusc.2016.02.229.Search in Google Scholar

[19] Zhu DQ, Yang CC, Pan J. New pyrometallurgical route for separation and recovery of Fe, Zn, In, Ga and S from jarosite residues. J Clean Prod. 2018;205:781–8. 10.1016/j.jclepro.2018.09.152.Search in Google Scholar

[20] Lv YN, Chen D, Tang JG. Recovery of iron from pyrite cinder with reduction-magnetic separation process in presence of sodium salt. Trans Indian Inst Met. 2019;72(2):501–10. 10.1007/s12666-018-1502-x.Search in Google Scholar

[21] Lei C, Yan B, Chen T, Xiao XM. Recovery of metals from the roasted lead-zinc tailings by magnetizing roasting followed by magnetic separation. J Clean Prod. 2017;158:73–80. 10.1016/j.jclepro.2017.04.164.Search in Google Scholar

[22] Jiao RM, Xing P, Wang CY. Recovery of iron from copper tailings via low-temperature direct reduction and magnetic separation: process optimization and mineralogical study. Int J Min Met Mater. 2017;24(9):974–82. 10.1007/s12613-017-1485-3.Search in Google Scholar

[23] Lei C, Yan B, Chen T, Quan SX, Xiao XM. Comprehensive utilization of lead–zinc tailings, part 1: pollution characteristics and resource recovery of sulfur. J Environ Chem Eng. 2015;3(2):862–9. 10.1016/j.jece.2015.03.015.Search in Google Scholar

[24] Rai S, Nimje MT, Chaddha MJ. Recovery of iron from bauxite residue using advanced separation techniques. Min Eng. 2019;134:222–31. 10.1016/j.mineng.2019.02.018.Search in Google Scholar

[25] Riesgo G, Elbaile L, Moriche R. Influence of the remnant magnetization, size distribution and content of soft magnetic reinforcement in micro-mechanical behavior of polymer matrix composites. Polym Test. 2019;79:106020. 10.1016/j.polymertesting.2019.106020.Search in Google Scholar

[26] Mahadevan S, Behera SP, Gnanaprakash G. Size distribution of magnetic iron oxide nanoparticles using Warren–Averbach XRD analysis. J Phys Chem Solids. 2012;73(7):867–72. 10.1016/j.jpcs.2012.02.017.Search in Google Scholar

[27] Lan YP, Liu QC, Meng F. Optimization of magnetic separation process for iron recovery from steel slag. J Iron Steel Res Int. 2017;24(2):165–70. 10.1016/S1006-706X(17)30023-7.Search in Google Scholar

[28] Li C, Sun HH, Yi ZL, Li LT. Innovative methodology for comprehensive utilization of iron ore tailings. J Hazard Mater. 2010;174(1–3):71–7. 10.1016/j.jhazmat.2009.09.019.Search in Google Scholar PubMed

[29] Bellam R, Anipindi NR, Jaganyi D. Effect of CTAB and SDS on base hydrolysis of iron(II)-sulphonated and unsulphonated phenyl-1,2,4-triazine complexes-akinetic and mechanistic study. J Mol Liq. 2018;258:57–65. 10.1016/j.molliq.2018.02.122.Search in Google Scholar

[30] Tripathy SK, Singh V, Murthy YR. Influence of process parameters of dry high intensity magnetic separators on separation of hematite. Int J Min Process. 2017;160:16–31. 10.1016/j.minpro.2017.01.007.Search in Google Scholar

[31] Jiang PG, Liu JS, Xiao YY. Recovery of iron from copper slag via modified roasting in CO–CO2 mixed gas and magnetic separation. J Iron Steel Res Int. 2020;27(7):796–806. 10.1007/s42243-020-00413-0.Search in Google Scholar

Received: 2020-10-21
Revised: 2020-12-01
Accepted: 2020-12-08
Published Online: 2021-02-10

© 2021 Bing Luo et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 29.3.2024 from https://www.degruyter.com/document/doi/10.1515/chem-2021-0194/html
Scroll to top button